Sunteți pe pagina 1din 30

Evaporative Characteristics of R-134a,

MP-39, and R-12 at Low Mass Fluxes


J. P. Wattelet, J. C. Chato, A. L. Souza, and B. R. Christoffersen
ACRCTR-35
For additional information:
Air Conditioning and Refrigeration Center
University of Illinois
Mechanical & Industrial Engineering Dept.
1206 West Green Street
Urbana,IL 61801
(217) 333-3115
May 1993
Prepared as part of ACRC Project 1
Refrigerant-Side Evaporation and Condensation Studies
J. C. Chato, Principal Investigator
The Air Conditioning and Refrigeration
Center was founded in 1988 with a grant
from the estate of Richard W. Kritzer, the
founder of Peerless of America Inc. A State
of Illinois Technology Challenge Grant
helped build the laboratory facilities. The
ACRC receives continuing support from the
Richard W. Kritzer Endowment and the
National Science Foundation. Thefollowing
organizations have also become sponsors of
the Center.
Acustar Division of Chrysler
Allied-Signal, Inc.
Amana Refrigeration, Inc.
Carrier Corporation
Caterpillar, Inc.
E. I. du Pont de Nemours & Co.
Electric Power Research Institute
Ford Motor Company
General Electric Company
Harrison Division of GM
ICI Americas, Inc.
Johnson Controls, Inc.
Modine Manufacturing Co.
Peerless of America, Inc.
Environmental Protection Agency
U. S. Anny CERL
Whirlpool Corporation
For additional information:
Air Conditioning & Refrigeration Center
Mechanical & Industrial Engineering Dept.
University of Illinois
1206 West Green Street
Urbana IL 61801
2173333115
ABSTRACT
EVAPORATIVE CHARACTERISTICS OF
R-134a, MP-39, AND R-12 AT LOW MASS FLUXES
J.P. Wattelet, J.C. Chata, A.L. Souza, B.R. Christoffersen
Air Conditioning and Refrigeration Center
Department of Mechanical and Industrial Engineering
University of Illinois at Urbana-Champaign
Experimental heat transfer coefficients for R-134a, MP-39, and R-12 are reported
herein. Tests were conducted using a single-tube evaporator test facility. The test section
used was a 2.43 m (8-foot) long, 7.04 mm (0.277-in) inside diameter, horizontal, smooth
copper tube. Heat was applied to the test section using electric resistance heaters. Test
parameters were varied as follows: mass flux, 25-100 kg/m
2
-s (19-76 klb
m
/ft
2
-hr); heat
flux, 2-10 kW/m2 (600-3200 Btu/hr-ft
2
); quality, 10-90 percent; saturation temperature,
-15 to 5C (5 to 41F). The heat transfer coefficients for R-134a were higher than those
of R-12 using equivalent mass flux and cooling capacity bases. The heat transfer
coefficients for MP-39 were lower than those of R-12 using equivalent mass flux and
cooling capacity bases. The correlations of Shah and Kandlikar were compared with the
experimental heat transfer coefficients. An empirical correlation was developed for the
wavy-stratified flow pattern occurring with the low mass fluxes. For correlating
purposes, some high mass flux data with annular flow patterns, reported previously, are
also included.
NOMENCLATURE
a exponent in Eq. (2)
A constant in Eq. (2)
As surface area
b exponent in Eq. (2)
B constant in Eq. (2)
Bo = ~ , boiling number
Gl
fg
D diameter
F = (Xcb , two-phase convection multiplier
(Xl
Fr
G
2
=-2-' Froude number
p gD
g gravitational acceleration
G mass flux
ifg
latent heat of vaporization
K
dxi
f
= ~ , Pierre boiling number
Lg
L tube length
M molecular weight
n exponent in Eq. (5), (7)
P
Pr
=-, reduced pressure
P crit
Ilc
Pr =_P , Prandtl number
k
q heat flux
R reduction parameter in Eq. (7)
ReI
GD(l- x), liquid Reynolds number
III
Relo
= GD , total liquid Reynolds number
III
2
S suppression factor
T s surface temperature
Tb bulk fluid temperature at saturation
x vapor quality
Xtt =c : x r ( :: r r lAlckhart-Martinelli parameter
Xu' =(
1
: x) 0.9 n, modified Lockhart-Martinelli parameter
Greek symbols
a heat transfer coefficient
( )
O.S( )0.1
n = o. transport property ratio parameter
p density
J.1 viscosity
A change in
Subscripts
b bulk fluid
cb convective boiling
crit critical
1 liquid phase
10 liquid only
nb nucleate boiling
s surface
TP two-phase
v vapor
3
INTRODUCTION
The prediction of heat transfer coefficients and pressure drops during forced-
convective evaporation is important for a wide range of processes, which include
applications relating to the power, chemical process, and refrigeration industries. The
first phase of extensive study of these heat transfer coefficients took place during the
1950s and 1960s. Appropriate mechanisms for forced-convective evaporation were
identified and dimensionless parameters representing these mechanisms were used in
developing correlations. However, magnitudes of resulting terms and transitions between
flow regimes were not extensively examined. Later, generalized correlations were
developed in the 1970s and 1980s from compiled databases of these earlier studies.
Improvements in predicting heat transfer coefficients were made, but the magnitudes of
each mechanism and transitions between these mechanisms were still not clearly defined.
Recently, because of the phaseout of chlorofluorocarbons outlined by the
Montreal Protocol and tightening energy efficiency standards imposed by the federal
government, the refrigeration industry, with the aid of chemical companies, has
identified alternative, ozone-safe refrigerants whose thermal and heat transfer properties
have not been determined. Experimental research to measure heat transfer coefficients of
these new chemicals has begun, including development of a test apparatus at the Air
Conditioning and Refrigeration Center. In addition to collecting baseline data for the
refrigerant industry, another opportunity has arisen to extend the existing databases
regarding evaporative heat transfer coefficients.
Most of the correlations developed in the literature are for annular flow situations
inside horizontal or vertical tubes. Smaller data bases have been generated regarding
wavy-stratified flows, which is the predominant flow pattern found in household
refrigerators and supermarket display case evaporators. These systems require much
4
smaller heat loads per tube than stationary air conditioning systems, and therefore need
much smaller mass flow rates.
A study was conducted to examine heat transfer coefficients of alternative
refrigerants for R-12 for low mass flux conditions. Experimental data was taken with
refrigerants R-134a and MP-39, alternative refrigerant replacements for R-12, and were
compared with baseline tests using R-12. Heat transfer coefficients were compared with
correlations of Kandlikar [1990] and Shah [1982]. An empirical correlation developed
for annular flow data using an asymptotic form was modified to account for the decrease
in heat transfer due to the wavy-stratified flow pattern in the low mass flux cases.
LITERATURE REVIEW
Initial methods were developed for convective boiling heat transfer along a liquid-
vapor interface. These methods were based on the premise that the mechanism of heat
transfer in forced convection was similar to single-phase forced convection [Chaddock
and Noerager, 1967]. By applying the Reynolds analogy that relates the energy transport
mechanism to momentum transport in convection, it was shown that the ratio between the
two-phase flow and the single-phase liquid heat transfer coefficients could be exclusively
correlated by the Lockhart-Martinelli parameter, Xtt. The form of this correlation was
(1)
It was also observed that nucleate boiling could occur simultaneously with
evaporation along an extensive liquid-vapor interface. Pierre[1956] developed a model
for the average heat transfer coefficient over a large quality change using Reynolds
number and a modified boiling number as follows:
(2)
To add nucleate boiling effects into the form of Eq. (1), Shrock and Grossman [1962]
introduced the boiling number, Bo, as follows:
5
a
TP
= F(Bo X )
, tt
0.
1
(3)
Rohsenow [1952] first proposed an additive model of the nucleate and convective
boiling heat transfer coefficients, and Chen [1966] utilized this based on the superposition
of heat transfer coefficients as follows:
(4)
where S is a suppression factor for nucleate boiling and F is the two-phase convective
multiplier for heat transfer as defined in the right hand side of Eq. (1). Recently, Jung
and Radermacher [1989] developed a correlation using this basic form.
Shah [1976] later utilized a different form by developing a generalized correlation
using several data bases from the literature which broke flow boiling into distinct regions:
a nucleate boiling dominated regime, a bubble suppression regime where nucleate boiling
and convective boiling are important, and a convective boiling dominated regime. The
correlation was in graphical form and was evaluated by taking the larger of the three heat
transfer coefficients calculated for the nucleate boiling, bubble suppression, and
convective boiling regions. The nucleate boiling term was characterized by the boiling
number while the convective boiling term was characterized by the convection number,
Co, which is a modified form of the Lockhart-Martinelli parameter. Shah [1982] later put
this correlation in equation form.
Recognizing that the boiling number could not accurately model the nucleate
boiling term alone, Kandlikar [1990] developed a similar correlation that multiplied the
boiling number by a fluid specific term which accounted for the different nucleate boiling
effects that occurred from fluid to fluid. This can be described as a "greater of the two"
(nucleate and convective boiling dominated) method.
The generalized correlations proposed by Shah and Kandlikar also account for the
loss of tube wetting in horizontal flow for low mass fluxes. For wavy-stratified flows,
part of the tube wall remains dry and there is a loss of convective heat transfer area,
6
thereby decreasing the heat transfer coefficient. In these correlations, this loss of tube
wetting is accounted for by introducing a Froude number based correction factor to the
convective term.
Connecting the Chen and Shah methods is the form Kutateladze [1961] proposed
which is an asymptotic, power-type addition model for the nucleate and convective
boiling components:
(5)
For n equal to 1, the form becomes that of Chen. As n approaches infinity, the form
becomes the "greater of the two" method similar to that of Shah. This method was used
by Churchill [1972] to correlate the transition between forced convection and natural
convection heat transfer. Bergles and Rohsenow [1964], Steiner and Taborek [1992], and
Liu and Winterton [1988] have developed correlations utilizing this form. Choices of n
have ranged from 2 to 3.
EXPERIMENTAL APPARATUS AND METHODS
Loop
Figure 1 is a schematic of the experimental test facility. A variable-speed gear
pump was used to circulate the refrigerant around the loop, eliminating the need for a
compressor and an expansion device found in vapor-compression refrigeration systems.
This allowed testing capabilities in a pure refrigerant environment without the influence
of the oil from the compressor. A 6-kW preheater was used to control inlet qualities to
the test section. The preheater was a 3-pass, horizontal, serpentine, copper coil wrapped
with strip heaters to provide heat input. Heat input rates to the preheater were set using
variacs which were manually controlled and were measured with a watt transducer. Heat
was removed from the refrigerant using two parallel condensers cooled by an ethylene
glycol-water mixture from a chiller. Flow rates were measured using a Coriolis-type
mass flow meter. Absolute pressure was measured at several locations using strain-gage
7
pressure transducers. Fluid temperatures were measured using calibrated, type T
thermocouples.
Test section
The horizontal test section was a 2.43 m (8 foot) long copper tube with a 7.04 mm
(0.277 in) inside diameter. Heat input was provided to the test section using variac-
controlled, surface-wrapped heaters. The heat rate was measured by a watt transducer.
To reduce heat gain from the environment, the test section was covered with 50 mm (2 in)
of foam insulation. Sixteen type T thermocouples were soldered in grooves cut
longitudinally along the test section to measure the surface temperatures. Figure 2 shows
the location of these thennocouples. Bulk fluid temperatures were measured at the inlet
ant outlet of the test section using type T thermocouple probes. Differential pressure
across the test section was measured using a pressure transducer across the pressure taps
shown in Fig. 2. Sight glasses with the same inside diameter as the test section were
installed at the inlet and outlet of the test section. Flow visualization was enhanced using
a strobe light.
Data Collection
Prior to running, the system loop was evacuated and then charged with refrigerant.
Testing showed that the saturation pressure based on thennocouple readings matched the
measured pressure within 2 kPa (0.3 psi). Thermocouples were calibrated in a
thermostatic bath over the testing temperature range and pressure transducers were
calibrated using a dead weight tester. The estimated uncertainty of the thermocouples
was O.15C (O.27 OF) and the estimated uncertainty of the pressure transducers was 2
kPa (0.3 psi). Uncertainty of the differential pressure transducers was estimated at 0.15
kPa (0.02 psid). Watt transducers were factory calibrated within 1O Wand this number
has been verified during single-phase energy balance testing. Heat gains to the test
section were also determined through single-phase energy balance testing.
8
Data collection was performed using a micro-computer and data acquisition
system. The data acquisition system consisted of 5 terminal panels for 40 transducer
connections, four data acquisition cards connected to the computer via Nubus slots, and
data acquisition software. Testing was conducted at steady-state conditions, which were
monitored and controlled by the above-mentioned system.
Parameters controlled during tests were mass flux, heat flux, inlet quality, and
saturation temperature. Steady-state conditions, reached in approximately 15 minutes to
two hours, were assumed when the time variation of saturation temperature was less than
O.IC (0.18P) for five minutes. The controlled parameters also had to be within the
following range of target values: mass flux, 5%; heat flux, 5%; saturation temperature,
a.5C. Once steady-state conditions were achieved, the transducer signals were logged
into a data acquisition output file. The channels were scanned once every second for a
period of 60 seconds. The inlet pressure to the test section was then recorded at 100 Hz
for 20 seconds. The data was then reduced using a spreadsheet macro. Refrigerant
transport and thermodynamic properties used during data collection and reduction were
obtained from McLinden et. al [1991].
Experimental Results and Uncertainties
Experimental heat transfer coefficients were determined by the convective law of
cooling using the circumferentially averaged values of surface temperatures, the linearly
interpolated values of bulk fluid temperature, the surface area of the test section, and the
heat input rate to the test section as follows:
h = q/ As
(Ts - T
b
)
(6)
Since surface temperatures were measured inside grooves located along the external
surface of the copper tube, the temperature drop across the tube wall needed to be
considered. However, its value was determined to be negligibly small, and, therefore, it
was disregarded. Axial heat conduction along the length of the tube was also neglected.
9
Heat gain from the environment was determined during single-phase energy balance
testing.
Uncertainties for the experimental heat transfer coefficients were determined
using the method of sequential perturbation as outlined by Moffat [1988]. Uncertainties
in each of the independent variables used to calculate the heat transfer coefficient from
Eq. (6) were estimated based on calibration and examination of system-sensor interaction
errors, system disturbance errors, and other errors. Uncertainties ranged from 3% to
12.5% for low mass flux testing.
ANNULAR FLOW DATA
Tests were previously conducted for high mass fluxes and heat fluxes similar to
those found in automotive air conditioning and stationary air conditioning evaporators.
Initial results were given in Wattelet et. al [1992] and the complete set of data is
summarized in Panek [1991]. The test section consisted 10.92 mm (0.43 in) inside
diameter, 2.43 m (8 foot) long copper tube. Test parameters for the high mass flux testing
are given in Table 1.
Table 1. High mass flux test parameters
Parameter Value
Saturation temperature -5 to 15C (23 to 59 OF)
Mass flux 200-500 kglm
2
-s (152-380
klbm/ft
2
-hr)
Heat flux 5-30 kW/m2 (1600 to 9600
Btulhr-ft
2
)
Quality 10-90 %
10
Refrigerants R-134a, R-12, MP-39 were used as test fluids. Flow patterns were
detennined by strobe-light enhanced visual observation through sight glasses at the inlet
and outlet of the test section. For higher mass flux testing, the predominant flow pattern
was annular flow.
Figure 3 shows the variation of heat transfer coefficient versus quality for annular
flow tests with a fixed mass flux and varying heat flux for R-134a, R-12, and MP-39. As
can be noted, for low heat fluxes the heat transfer coefficient increases with quality.
Intense evaporation at the liquid-vapor interface diminishes the liquid film thickness,
reducing the thennal resistance, which is associated with heat conduction across the film.
Nucleate boiling appears to be largely suppressed for these low heat flux cases. As heat
flux increases, the heat transfer coefficients increase in the lower quality region and
eventually merge at higher qualities with the heat transfer coefficients for low heat flux
cases. Nucleate boiling at these lower qualities enhances the heat transfer coefficient. At
higher qualities, nucleate boiling is again largely suppressed due to significant surface
cooling promoted by the thinning of the annular film.
R-134a and R-12, both single-component fluids, exhibit similar behavior
qualitatively. For an equivalent, mass flux, heat flux, saturation temperature, and quality,
R-134a had approximately a 25% higher heat transfer coefficient on average than R-12
for annular flow tests. On an equivalent cooling capacity basis, R -134a had
approximately a 2% higher heat transfer coefficient on average than R-12 for annular
flow tests.
MP-39, a 52% / 33% /15% near azeotropic mixture of R-22 / R-124 / R-152a, has
similar convective heat transfer properties to R-134a. For low heat flux tests shown in
Fig. 3, the heat transfer coefficients for MP-39 are roughly the same as R-134a.
However, its nucleate boiling characteristics are significantly diminished compared with
R-134a and R-12. For higher heat flux testing where nucleate boiling effects are more
important, R-134a has a higher heat transfer coefficient than MP-39. Overall, MP-39 has
11
a 10% higher heat transfer coefficient on average compared with R-12 for annular flow
tests. As will be noted in the next section, this is quite different than the comparisons
between MP-39 and R-12 for low mass flux tests with wavy-stratified flow patterns.
STRATIFIED FLOW DATA
Test parameters for the low mass flux testing are given in Table 2. Refrigerants
R-134a, R-12, MP-39 were used as test fluids. For low mass flux testing, the
predominant flow pattern was wavy-stratified flow.
Table 2. Low mass flux test parameters
Parameter Value
Saturation temperature -15 to 5C (5 to 41F)
Mass flux 25-100 kg/m
2
-s (19-76
klbm/ft
2
-hr)
Heat flux 2-10 kW/m2 (600 to 3200
Btulhr-ft
2
)
Quality 10-90 %
For wavy-stratified tests, there is no major effect of quality on the
circumferentially averaged heat transfer coefficients, as can be seen in Fig. 4 for R-134a,
R-12, and MP-39. However, as the heat flux increases, the heat transfer coefficient also
increases. Compared with the results in the annular flow regime, convective boiling is
diminished while nucleate boiling does not appear to be suppressed at higher qualities or
lower heat fluxes. The decrease in convective boiling can be attributed to the reduction in
available surface area for convective boiling, a decrease in turbulence due to the
decreased Reynolds numbers for the low mass fluxes, and a decrease in slip ratio between
the vapor and liquid streams.
12
R-134a again has a higher heat transfer coefficient than R-12 based on equivalent
mass flux and cooling capacity comparisons for wavy-stratified tests. However, MP-39
shows a significant decrease in heat transfer coefficient for low mass flux cases. Figure 5
is a comparison of average heat transfer coefficients over the quality range of 20-90
percent for R-134a, R-12, and MP-39. For high mass fluxes, MP-39 has comparable heat
transfer coefficients to R-134a. As mass flux is decreased, the heat transfer coefficients
for MP-39 fall below R-12 by a significant amount. For the wavy-stratified cases, MP-39
has a 20% lower heat transfer coefficient on average compared with R -12. In addition to
the decrease in nucleate boiling due to concentration gradients in the liquid, the lower slip
ratios and turbulence in the wavy-stratified flows also seem to decrease the mixing in the
vapor stream and concentration gradients may also be set up in the vapor.
CORRELATION COMPARISON
Two heat transfer correlations were selected from the literature to compare with
the experimental values of heat transfer coefficient. The correlations of Shah [1982] and
Kandlikar [1990] were selected because of their use of R-12 as a test fluid and the
presence of a Froude number dependent term to account for wavy-stratified flow. The
Shah correlation was the first generalized correlation and evaluates equations for the
nucleate boiling, bubble suppression, and convective boiling dominated regimes and
takes the largest of the three values. The Kandlikar correlation uses the greater of the
nucleate boiling and convective boiling dominated forms. It differs from the Shah
correlation in its use of boiling number for both the nucleate boiling and convective
boiling regimes. It also uses a fluid specific term to account for variations of the nucleate
boiling component of each form.
Figures 6, and 7 are plots of the predicted heat transfer coefficients for R-12 and
R-134a from the correlations of Shah and Kandlikar, respectively, versus experimental
heat transfer coefficients for R-12 and R-134a for annular and wavy-stratified flow. The
13
mean deviation of the three correlations are given in Table 3 for both wavy-stratified and
annular flow tests. The Kandlikar correlation does a slightly better job predicting the data
than the Shah correlation.
Table 3. Comparison of mean deviation between the various correlations and the present
experimental data for R-134a and R-12
Correlation R-134a Annular R-134a Wavy R-12 Annular R-12 Wavy
Kandlikar 9.9% 12.2% 9.7% 14.0%
Shah 11.8% 20.1% 16.4% 16.9%
Wattelet et. al 7.6% 6.6% 6.8% 7.1%
Figure 8 is a plot of the predicted heat transfer coefficients of MP-39 from the
correlation of Shah with the experimental heat transfer coefficients of MP-39. Because of
the reduction of both nucleate boiling and convective boiling heat transfer for the
zeotrope, the unmodified Shah correlation severely overestimates by more than 50% the
experimental heat transfer coefficients in the wavy-stratified flow regime. However, for
annular flows, the convective correlating form does an adequate job of predicting the MP-
39 heat transfer coefficients. As was discussed earlier, nucleate boiling effects are
dramatically reduced for zeotropic mixtures. To properly correlate zeotropic mixtures,
modifications will need to be made to the existing correlations in the literature, modifying
the nucleate boiling term for annular flow and both the nucleate boiling and convective
boiling terms for wavy-stratified flow. Although not discussed in this paper, Kandlikar
proposed an initial modification to his generalized correlation for exactly this purpose for
binary mixtures [1991]. These modifications to the Kandlikar correlation among others
will become even more important as many zeotropic mixtures are currently the leading
candidates for replacement of R-22.
14
MODEL SELECTION AND COMPARISON
The superposition model, the "greater of the two" model, and the asymptotic
model were all discussed earlier in the literature review. After extensive evaluation of
these forms, the asymptotic model was chosen to be the best form to correlate the
experimental data. Correlations from nucleate pool boiling can be used for the nucleate
boiling term while a convective form similar to the proposed form of Chen [1966] can be
evaluated experimentally and used for the convective boiling term.
The main feature of this form is the "built in" suppression of the weaker
component. Table 4 shows an example of this with n equal to two for the asymptotic
form. For a large convective component and a small nucleate boiling component, the
total two-phase heat transfer coefficient is made up almost entirely of the convective
boiling component. For a mixed situation where both nucleate boiling and convective
boiling both occur, the total two-phase heat transfer coefficient is made up of a
combination of the two components. For a nucleate boiling dominated situation, the total
two-phase heat transfer coefficient is made up almost entirely by the nucleate boiling
component.
Table 4. Asymptotic form examples
h (nucleate boiling) h (conv. boiling) h (two-phase)
5000 1000 5036
3000 3000 3959
1000 5000 5036
Using annular flow data for R-134a and R-12 discussed above, an asymptotic
correlation was developed using the following equations:
[
n n ]l/n 25
(XTP = (Xnb + (Xcb n = .
(7)
15
a
nb
=
acb = FaiR
F = 1 + 1.
k
a
l
= 0.023-
1

D
R
_ GD(l-x)
e
l
-
f..11
(
1_X)009
X = - Q
u
x
Q =
Pr = f..1l
C
Pl
I k
I
R = ifFrl < 0.25
R = 1 ifFrl 0.25
To account for the decease in convective heat transfer due to loss in convective
boiling surface area and a loss of turbulence for lower Reynolds number flows, a Froude
dependence has been added to the convective term. Compared with the Shah and
Kandlikar correlations, this term begins to take effect at a much higher Reynolds number.
This offsets the overestimation of the single-phase liquid heat transfer coefficient through
use of the Dittus-Boelter correlation [McAdams, 1942] for tests with Reynolds numbers
below 10,000. Kandlikar [1991] and others have recently suggested using the Gnielinski
correlation [1976] for test conditions with Reynolds numbers below 10,000. However,
the Gnielinski correlation only works for Reynolds numbers above 4,000 and is not
suggested for extrapolation below this value because of the presence of a (Re-1000) term
in the correlation. Many practical uses of refrigerants inside horizontal tubes, such as in
household refrigerator evaporators, have Reynolds numbers below 4,000. Because the
form of the Dittus-Boelter correlation is more tractable to modification, the Dittus-Boelter
correlation was selected for use in the convective boiling term of Eq. (7).
Several nucleate pool boiling correlations have been developed over the past
several decades. Initial correlations suggested by Chen [1966], such as the Forster-
16
Zuber correlation [1955] for nucleate pool boiling, are difficult to calculate and require
knowledge of data such as surface temperatures and surface tension, which are not always
available. This lack of surface tension data may become even more important for
correlating heat transfer coefficients for zeotropic mixtures. Two recent pool boiling
correlations have been developed that are more accurate than some of the original
correlations and are easier to evaluate. The Cooper correlation [1984] is based on
reduced pressure, heat flux, and molecular weight and is of the same order of accuracy as
the Forster-Zuber correlation, but is much easier to evaluate. The other recent correlation
developed in the literature is that of Stephan and Abdelsalam[1980] , and is used in the
Jung-Radermacher correlation [1989]. Insufficient data has been collected in the present
study for nucleate boiling dominated regimes of flow boiling to adequately distinguish
between the two. Because of model simplicity, need for surface tension in the Stephan-
Adelsalam correlation, and similar accuracy in correlating the present flow boiling data,
the Cooper correlation was selected for the nucleate boiling term in Eq. (7).
A modified form of the convective term in the Chen correlation was selected.
Kenning and Cooper [1989] have shown this to be the appropriate form for this term.
However, the Chen correlation has been found to underestimate their data and others.
The convective boiling dominated experimental data in this paper also is underestimated
by the Chen correlation. The form for the two-phase multiplier, F, in Eq. (7) is
approximately 10 to 30% higher than the Chen two-phase multiplier between qualities of
10 and 90 percent for refrigerants R -134a and R -12.
The value of n selected in the asymptotic model was 2.5. This values was
determined by a regression analysis for values of n between 1 to 3.
Figure 9 is a comparison of the experimental values of heat transfer coefficient for
annular and wavy-stratified flow data for R-134a and R-12 and the predicted values of
Eq. (7). The mean deviation of the correlation for the experimental data is shown in
Table 3. These results compare very favorably to those of the other correlations shown
17
in Table 3 and should give much credibility to the asymptotic form used in Eq. (7). MP-
39 heat transfer coefficient values, however, are also severely overestimated by Eq. (7),
mainly due to the stronger suppression of nucleate boiling for the zeotropic mixture. Use
of the convective boiling term only results in a mean deviation similar to that of Shah's
correlation for the annular flow data. Further testing and analysis needs to be performed
before corrections to the nucleate boiling and convective boiling terms of Eq. (7) can be
made.
CONCLUSIONS
Experimental heat transfer coefficients for R-12, R-134a, and MP-39 were
determined for both annular flow and wavy-stratified flow. For equivalent mass flux,
heat flux, saturation temperature, and quality, the heat transfer coefficients for R-134a
were on average 25% higher than those of R-12. MP-39 heat transfer coefficients for
similar conditions were 10% higher on average than R-12 for annular flow tests and 20%
lower on average than R-12 for wavy-stratified tests. For equivalent cooling capacities,
R-134a had slightly higher heat transfer coefficients than R-12 and MP-39 had lower heat
transfer coefficients than R-12 for both wavy-stratified and annular flows.
Heat transfer coefficient correlations from the literature predicted heat transfer
coefficients for low mass flux conditions for R-134a and R-12 with mean deviations of
under 20%. MP-39 experimental heat transfer coefficients were overestimated by all
correlations for wavy-stratified flows. Significant reductions in both nucleate and
convective boiling due to concentration gradients in the liquid and vapor streams and the
lack of turbulent mixing in the wavy-stratified flow contribute to this degradation from
the predicted values using the weighted averages of liquid and vapor properties.
A new heat transfer coefficient correlation was developed using an asymptotic
form. The Cooper correlation was selected for the nucleate boiling term. A two-phase
multiplier, F, was developed using a modified Lockhart-Martinelli parameter for the
18
convective term. For wavy-stratified flow, a Froude number correction was added to the
convective boiling term to account for the decrease in available convective heat transfer
area and the decrease in turbulence as predicted by the Dittus-Boelter single-phase heat
transfer correlation. This correlation predicted both the annular and wavy-stratified flow
data well.
ACKNOWLEDGMENTS
The support by the Air Conditioning and Refrigeration Center of the University of
Illinois is gratefully acknowledged by the authors. They would also like to extend their
appreciation to E.I. du Pont de Numours and Company for providing the refrigerants
used in this investigation, and particularly to Dr. Don Bivens.
19
REFERENCES
Bergles, A.E., and W.M. Rohsenow. 1964. The determination of forced convection
surface boiling. ASME Journal of Heat Transfer 86: 365-372.
Chaddock, J.B., and J.A. Noerager. 1966. Evaporation of refrigerant 12 in a horizontal
tube with constant wall heat flux. ASH RAE Transactions 72(1): 90-103.
Chen, J.C. 1966. A correlation for boiling heat transfer to saturated fluids in convective
flow. Industrial and Engineering Chemistry, Process Design and Development
5(3): 322-329.
Churchill, S.W., and R. Ugasi. 1972. A general expression for the correlation of rates of
transfer and other phenomena. Ai. Ch. E. Journal 18(6): 1121-1128.
Cooper, M.G. 1984. Saturation pool boiling--A simple correlation. International
Chemical Engineering Symposium Series 86:785-792.
Eckels, S.J., and M.B. Pate. 1990. An experimental comparison of evaporation and
condensation heat transfer coefficients for HFC-134a and CFC-12. International
Journal of Refrigeration 14(2): 70-77.
Forster, H.K., and N. Zuber. 1955. Dynamics of vapour bubbles and boiling heat
transfer. A. I. Ch. E. Journal 1: 531-535.
Gnielinski, V. 1976. New equations for heat and mass transfer in turbulent pipe and
channel flow. International Chemical Engineer 16: 359-368.
Jung, D.S., and R. Radermacher. 1989. A study of flow boiling heat transfer with
refrigerant mixtures. International Journal of Heat and Mass Transfer 32(9):
1751-1764.
Kandlikar, S.G. 1990. A general correlation for saturated two-phase flow boiling heat
transfer inside horizontal and vertical tubes. ASME Journal of Heat Transfer.
112: 219-228.
Kandlikar, S.G. 1991. Correlating flow boiling heat transfer data in binary systems.
Phase Change Heat Transfer --ASME HTD 159: 163-170.
Kenning, D.B.R., and M.G. Cooper. 1989. Saturated flow boiling of water in vertical
tubes. International Journal of Heat and Mass Transfer 32(3): 445-458.
Kutateladze, S.S. 1961. Boiling heat transfer. International Journal of Heat and Mass
Transfer 4: 31-45.
Liu, Z., and R.H.S. Winterton. 1988. Wet wall flow boiling correlation with explicit
nucleate boiling term. Proceedings of 5th Int. Symp. Multiphase Transport and
Particulate Phenomena, Dec. 12-14, Miami Beach, FI.
McAdams, W.H. 1942. Heat transmission. 2d ed. New York: McGraw-Hill.
McLinden, M., J. Gallagher, and M. Huber. 1991. REFPROP 3.0. NIST
Thermodynamic Properties of Refrigerants and Refrigerant Mixture Database.
U.S. Department of Commerce. National Institute of Standards and Technology.
20
Moffat, R.J. 1988. Describing the uncertainties in experimental results. Experimental
Thermal and Fluid Science 1: 3-17.
Panek, J.S. 1991. Evaporation heat transfer and pressure drop in ozone-safe
refrigerants and refrigerant-oil mixtures. Master of Science Thesis. University
of Illinois at Urbana-Champaign.
Pierre, B. 1956. Coefficient of heat transfer for boiling freon-12 in horizontal tubes.
Heat ing and Air Treatment Engineer 19: 302-310.
Rohsenow, W.M. 1952. A method of correlating heat transfer data for surface boiling of
liquids. ASME Transactions 74: 969-976.
Shah, M.M. 1976. A new correlation for heat transfer during boiling through pipes.
ASHRAE Transactions 82(2): 66-86.
Shah, M.M. 1982. Chart correlation for saturated boiling heat transfer: Equations and
further study. ASHRAE Transactions 88(1): 185-196.
Shrock, V.E., and L.M. Grossman. 1962. Forced convection boiling in tubes. Nuclear
Science and Engineering 12(4): 474-481.
Steiner, D., and J. Taborek. 1992. Flow boiling heat transfer of single components in
vertical tubes. Heat Transfer Engineering pp. 43-69.
Stephan, K., and M. Abdelsalam. 1980. Heat transfer correlation for natural convection
boiling. International Journal of Heat and Mass Transfer 23: 73-80.
Wattelet, J.P. et. al. 1992. Experimental evaluation of convective boiling of refrigerants
HFC-134a and CFC-12. Symposium on Basic Aspects of Two-Phase Flow and
Heat Transfer /V--Heat Transfer in Mixtures and Refrigerants, presented at the
29th National Heat Transfer Conference, San Diego, California.
21

Preheater
T
Test Section
Flow Meter Pump
Pressure sensor Differential pressure
Valve
.. Flow direction
(!) Temperature sensor
c:J Sight glass
Figure 1. Refrigerant Loop

QJ OJ
I J t

m Pressure tap
Section AA:
e
t

Thermocouple bead
Figure 2. Evaporation Test Section
22
8000
f 6000
i
'II
~
4000
0
!
2000
:
J:
0
0 0.2 0.4 0.6 0.8
Quality
(a)
8000
0
q.SkW/rr1l
q.l0 kW/rrl
6000 6
q.20 kW/rrl

q.30 k W / ~

4000
2000
" A.. Q..O
< .. 6 . ~ . '0
A 6. 6 _t 6. V'
.. .. AA 6. 6. t <tie 'V
. ~ .
'i <811.
'UP
o
0.2 0.4 0.6 0.8
OJaiity
(b)
8000
0 q_SkW/m2
q.l0kW/m2
6000 6
q.20kW/m
2
q.30kW/m2
4000
. .... '\"
M A." ". -: \. !_. .."
4" _ "tI._
-., .. ~
c!illl"
2000
o
0.2 0.4 0.6 0.8
Quality
(c)
Figure 3. Heat transfer coefficient versus quality during annular flow for (a) R-134a; (b)
R-12; (c) MP-39. Mass flux: 300 kg/m
2
-s; Saturation temperature: 5C; Tube
diameter: 10.92 mm.
23
1500
JZ 1200
[
I
I
i
if
E
~
. j
]
I
I!
I-
I
if
~
i
1I
~
8
I
:
:z:
800
600
300
0
0.2
1500
1200
1100
800
300
0
0.2
1500
1200
900
600
300
0
..
A
A
A



0
0
0 00
0
0 0 0
00 0
0
q-2 kW/rJ
q-3kW/rI-
A
q..s kW/rJ
0.3 0.4 0.5 0.6 0.7 0.8 0.9
Quality
(a)
..
..
A
..

0
0 0 0
0
0

0
0
0
0 0
q-2 kW/rfI.
0
q_3 kW/m
2
..
q..s kW/rJ
0.3 0.4 0.5 0.6 0.7 0.8 0.9
OJaiity
(b)
0
q_2kwlrfl.
q_3 kW/rJ
.. q_S kW/rI-
.. ..
0
0
0 0
0
0
0.2 0.3 0.4 0.5 0.6 0.7 0.8 0.9
OuaIRy
(c)
Figure 4. Heat transfer coefficient versus quality during wavy-stratified flow for (a) R-
134a; (b) R-12; (c) MP-39. Mass flux: 50 kg/m
2
-s; Saturation temperature: 5C;
Tube diameter: 10.92 mm.
24
4000
~
3500
~
3000
i
2500

2000
~ 1500
~

R-12
"
R-134a
0 MP-39
~
"

D

"
~
i
1000
:I:
500
" 8
"

0
D
0
o 50 100 150 200 250 300 350
Figure 5. Heat transfer coefficient versus mass flux for R-134a, R-12, and MP-39. Heat
flux: 5 kW/m2; Saturation temperature: 5C; Tube diameter, 7.04 mm.
6000
D R-12 wavy-stratified
5000 0 R-134a wavy-stratnied
q
"
R-12 annular
1:

R-134a annular
~
4000
.<::
to
.<::
3000 C/)
i
li
j
2000
"-
.<::
1000
0
o 1000 2000 3000 4000 5000 6000
h Experimental (W/ni!K)
Figure 6. Predicted heat transfer coefficient using the Shah correlation versus
experimental heat transfer coefficient for R-134a and R-12 during wavy-stratified and
annular flows.
6000
o R-12 wavy-stratified
~
5000
D R-134a wavy-stratified
" R-12 annular
~
4000
12
R-134a annular
'6
c:
3000
~
i
.'il
2000
i
"-
.<::
1000
0
0 1000 2000 3000 4000 5000 6000
hExperimentai (W/rJk)
Figure 7. Predicted heat transfer coefficient using the Kandlikar correlation versus
experimental heat transfer coefficient for R-134a and R-12 during wavy-stratified and
annular flows.
25
1600
1400
SZ
1200
'E
~
1000
I
BOO
)
BOO
~
Q.
400
z::;
200
0
5000
SZ
4000
'E
!
3000
'Iii
6j
) 2000
~
Q.
z::;
1000
0
0
0
o MP-39 wavy-stl'lllilled I
200 400 600 BOO 1000 1200 1400 1600
h Experimental (W/m
2
K)
(a)
o Mp39 wavy-stl'lllilled
1000 2000 3000 4000 5000
h Experimental (W/rifK)
(b)
Figure 8. Predicted heat transfer coefficient using the Shah correlation versus the
experimental heat transfer coefficient for MP-39 during wavy-stratified and annular
flows.
26
6000
S2"
5000
"E
!
iii
1&
4000
I
3000
!
2000
~
Q.
1000
.<:
o R-12 wavy-stratified
" R-134a wavy-stratified
'" R-12 annular
o R-134a annular
0
0 1000 2000 3000 4000 5000 6000
h Experlmental(w/m2K)
Figure 9. Predicted heat transfer coefficient using the Wattelet et. a1 correlation versus
the experimental heat transfer coefficient for R-134a and R-12 during wavy-stratified and
annular flows.
27

S-ar putea să vă placă și